hemorheological behavior of glucose-rich rbc suspensions

22
In vitro particulate analogue fluids for experimental studies of rheological and hemorheological behavior of glucose-rich RBC suspensions Diana Pinho, Laura Campo-Deaño, Rui Lima, and Fernando T. Pinho Citation: Biomicrofluidics 11, 054105 (2017); doi: 10.1063/1.4998190 View online: http://dx.doi.org/10.1063/1.4998190 View Table of Contents: http://aip.scitation.org/toc/bmf/11/5 Published by the American Institute of Physics

Upload: others

Post on 28-Mar-2022

0 views

Category:

Documents


0 download

TRANSCRIPT

In vitro particulate analogue fluids for experimental studies of rheological and hemorheological behavior of glucose-rich RBC suspensions Diana Pinho, Laura Campo-Deaño, Rui Lima, and Fernando T. Pinho
Citation: Biomicrofluidics 11, 054105 (2017); doi: 10.1063/1.4998190 View online: http://dx.doi.org/10.1063/1.4998190 View Table of Contents: http://aip.scitation.org/toc/bmf/11/5 Published by the American Institute of Physics
1ESTiG, Polytechnic Institute of Braganca, C. Sta Apolonia, 5301-857 Braganca, Portugal 2CEFT, DEMec, Faculty of Engineering, University of Porto, Rua Dr. Roberto Frias, 4200-465 Porto, Portugal 3MEtRiCS, Mechanical Engineering Department, University of Minho, Campus de Azurem, 4800-058 Guimar~aes, Portugal
(Received 11 March 2017; accepted 28 July 2017; published online 21 September 2017)
Suspensions of healthy and pathological red blood cells (RBC) flowing in
microfluidic devices are frequently used to perform in vitro blood experiments for
a better understanding of human microcirculation hemodynamic phenomena. This
work reports the development of particulate viscoelastic analogue fluids able to
mimic the rheological and hemorheological behavior of pathological RBC
suspensions flowing in microfluidic systems. The pathological RBCs were obtained
by an incubation of healthy RBCs at a high concentration of glucose, representing
the pathological stage of hyperglycaemia in diabetic complications, and analyses of
their deformability and aggregation were carried out. Overall, the developed
in vitro analogue fluids were composed of a suspension of semi-rigid microbeads in
a carrier viscoelastic fluid made of dextran 40 and xanthan gum. All suspensions of
healthy and pathological RBCs, as well as their particulate analogue fluids, were
extensively characterized in steady shear flow, as well as in small and large ampli-
tude oscillatory shear flow. In addition, the well-known cell-free layer (CFL) phe-
nomenon occurring in microchannels was investigated in detail to provide
comparisons between healthy and pathological in vitro RBC suspensions and their
corresponding analogue fluids at different volume concentrations (5% and 20%).
The experimental results have shown a similar rheological behavior between the
samples containing a suspension of pathological RBCs and the proposed analogue
fluids. Moreover, this work shows that the particulate in vitro analogue fluids used
have the ability to mimic well the CFL phenomenon occurring downstream of a
microchannel contraction for pathological RBC suspensions. The proposed particu-
late fluids provide a more realistic behavior of the flow properties of suspended
RBCs when compared with existing non-particulate blood analogues, and conse-
quently, they are advantageous for detailed investigations of microcirculation.
Published by AIP Publishing. [http://dx.doi.org/10.1063/1.4998190]
I. INTRODUCTION
The behavior of separated, washed, and resuspended red blood cells (RBC), also called
erythrocytes, flowing through microfluidic devices has been studied for several years. This
blood element is the most concentrated component in blood and is the main cause of its non-
Newtonian behavior at low shear rates, where the blood viscosity is high due to the formation
of RBC agglomerates known as rouleaux. At high shear rates, the disruption of the agglomer-
ates and consequent alignment of the deformable cells lead to a reduction of the viscosity
(shear-thinning). The orientation of the RBCs at high shear rates and their tendency to migrate
to the center line of the microchannels (in vitro)1–3 or microvessels (in vivo)4,5 originate the for-
mation of a cell depleted layer near the walls, known as the cell-free layer (CFL). The CFL is
influenced by the formation of aggregates, cell interactions and deformability, hematocrit, flow
1932-1058/2017/11(5)/054105/21/$30.00 Published by AIP Publishing.11, 054105-1
BIOMICROFLUIDICS 11, 054105 (2017)
rate, viscosity, and geometry6,7 and is a microscopic level phenomenon that occurs in microflui-
dic devices2,6,8,9 and in microvessels10–12 with dimensions in the range of 300 lm down to
10 lm. Several studies have been performed both in vivo11,13,14 and in vitro2,9,15–18 to better
understand the CFL formation, which influences its thickness and the advantages and disadvan-
tages of the presence of this layer in the human microcirculatory system6 and in microchan-
nels.8,16,19 It is also important to refer that in real blood, there is a migration of white blood
cells from the core towards the wall region of the vessels, a phenomenon called margin-
ation,20,21 but this issue is outside the scope of this work. Additionally, it is also known that
CFL is influenced by the presence of several pathological disorders that also change the rheo-
logical properties of whole blood22,23 as is the case of diabetic disorders.
Diabetes mellitus is a metabolic disorder that reduces the effects of insulin on the human
body because either the pancreas is unable to produce enough insulin (type 1 diabetes) or the
body is resistant to the effects of insulin and in addition it does not produce enough insulin to
maintain a normal glucose level (type II diabetes). As a result, when the glucose concentration
increases in the bloodstream, or hyperglycemia, we have a condition that can affect people with
both type I and type II diabetes. The high concentration of glucose in plasma causes alterations
in the lipid-protein interactions and consequent glycation of erythrocyte protein membranes,
resulting in a decrease in RBC deformability and changes in their shape, leading to the forma-
tion of flat or discocyte cells. As a consequence, there is an increase in the internal viscosity of
RBCs due to structural alterations in the hemoglobin molecule and also an enhancement of
aggregate formation that increases the blood viscosity.24,25 Hyperglycemia is the most important
factor on the onset, and the progress of diabetic complications and untreated hyperglycemia can
cause long-term complications.
Hyperglycemia in diabetes was observed in several studies where the reduction in the RBC
deformability and the increase in the aggregation level were found to promote the enhancement
of the CFL thickness.22,23,26 The most frequently used in vitro experiment to study this patho-
logical disorder involves the incubation of healthy RBCs in high glucose concentrations as a
model to mimic in vivo conditions of hyperglycemia in diabetes. This approach has been suc-
cessfully used for a better understanding of RBC mechanical modifications and in particular the
investigation of the impairment of microcirculation in diabetic patients.22,23,27–29 In this work,
we have also applied this method to obtain pathological suspensions of RBCs due to the diffi-
culty in acquiring and handling pathological blood. Hence, it is crucial to develop blood ana-
logues that are able to mimic the hemodynamic and hemorheological properties not only of
healthy but also of the pathological blood behavior. Mimicking pathophysiological behavior of
the RBCs will be helpful to understand several pathological phenomena since working with
blood, and in particular, with pathological blood, continues to raise ethical, economical, and
safety issues. To the best of our knowledge, this is the first attempt to develop two-phase
in vitro analogue fluids that mimic the behavior of RBC suspensions at in vitro conditions
equivalent to the pathological in vivo stage of hyperglycaemia, a condition that occurs at dia-
betic complications.
Several authors have already proposed blood analogues that have a good agreement with
whole blood in terms of rheological properties, especially under steady shear conditions. Such
studies were performed with Newtonian blood analogues composed of mixtures of glycerol and
water,30–34 and others have used non-Newtonian fluids composed of an aqueous solution of xan-
than gum (XG) or/and polyacrylamide (PAA), diluted in water or/and glycerine.35,36 Campo-
Dea~no et al.37 have successfully developed whole blood analogues based on sucrose and dime-
thylsulfoxide with XG having viscosity curves and viscoelastic moduli similar to whole human
blood and in addition demonstrated that these analogues are suitable to be used in polydyme-
thylsiloxane (PDMS) microfluidic devices due to their refractive index matching. However,
in vitro blood flow experiments have shown that it is crucial to take into account the blood ele-
ments,1,2,15 in particular when in vitro flow studies intend to investigate microcirculation phe-
nomena occurring at hematocrits of about 20%–25%. For instance, Lima and co-workers1,2
have visualized and measured cell–cell interactions, and they have shown that RBC radial dis-
persion tends to increase with cell concentration and may influence the blood mass transport
054105-2 Pinho et al. Biomicrofluidics 11, 054105 (2017)
mechanisms. Hence, it is important hemodynamic flow phenomena that occur at the microscale
also be mimicked by particulate blood analogue fluids.
To date, there are very few works where particulate blood analogue fluids, i.e., a base fluid
containing solid suspended elements, have been proposed, and none of those are capable of
mimicking the blood element behavior. Maruyama et al.38,39 have developed a blood analogue
made of a Newtonian solvent containing a suspension of microcapsules to evaluate the absolute
hemolytic properties of centrifugal blood pumps. Later, Nguyen et al.40 have performed a simi-
lar study, although this time by using a non-Newtonian solvent. These latter studies present par-
ticulate blood analogues that are able to reproduce well the steady viscosity, but they did not
study any microscale flow phenomenon. An exception is the preliminary study by Calejo
et al.41 where constant and non-Newtonian continuous phases with suspended particles were
used to perform CFL flow studies in a PDMS microchannel. Their results have shown that the
CFL originated from the particles depends strongly on the base fluid, but otherwise further
developments and improvements are needed to mimic blood flow phenomena occurring at the
microcirculation level.
The main objective of this work is to develop in vitro analogue fluids able to mimic the
rheological properties of suspended pathological in vitro RBCs and also hemorheological flow
phenomena occurring in PDMS microchannels such as the CFL formation downstream of a
contraction. The proposed particulate analogue fluids are composed of a carrier fluid made of
115 ppm (w/w) of XG diluted in dextran 40 and with suspended semi-rigid microbeads with a
diameter of 10 lm, a dimension close to the size of human RBCs, at weight concentrations of
5% and 20%. Non-Newtonian properties of the developed analogues and of the RBC suspen-
sions were investigated by performing both steady and oscillatory shear flow tests.
Additionally, to analyse the ability of the proposed analogue fluids to mimic the flow behavior
of pathological in vitro RBCs suspensions, flow visualizations were performed in microchan-
nels. The CFL measurements obtained for the suspension of RBCs were compared with the
CFL thickness obtained by the corresponding analogue fluid.
The remainder of this paper is organized as follows: Section II comprises several subsec-
tions to explain the experimental framework around blood samples, analogue fluids prepara-
tion, and the setups used to acquire the rheological data and to perform the flow visualiza-
tions. Section III presents and discusses the results, and the main conclusions are provided in
Sec. IV.
Whole blood samples were collected into tubes with an anticoagulant (ethylenediaminete-
traacetic acid) from a healthy human donor, and the RBCs were separated from the plasma and
buffy coat by centrifugation. Then, RBCs were resuspended and washed twice with phosphate
buffer saline (PBS). To obtain the healthy fluids with the desired hematocrit (Hct) of 5% and
20% by volume, after washing, the RBCs were suspended in dextran 40 (Dx 40). Dextran 40
has a density of 1042 kg/m3 and a constant viscosity of 5.2 mPa.s at 20 C and is commonly
used in experimental studies as a substitute of blood plasma (q¼ 1021 kg/m3) since it mini-
mizes the sedimentation of blood cells during the experiments and reduces cell clogging.19 The
experiments of the blood flow and measurements of blood rheology were immediately per-
formed after blood preparation.
To obtain the pathological RBCs, we have performed a twelve hour incubation of the washed
erythrocytes in a highly concentrated glucose solution at 37 C. The glucose incubation medium
was prepared by diluting 100 mM of glucose in PBS. A normal blood glucose level in non-
diabetic humans (as our healthy donor), before a meal, is about 5 mM, so 100 mM represents a
very high glucose concentration.25,28 After the incubation time, the RBCs were washed twice
with PBS in order to remove any remaining traces of the glucose-rich medium. Finally, we resus-
pended the glucose-rich RBCs at the desired Hct in Dx 40 to perform all the experiments.
054105-3 Pinho et al. Biomicrofluidics 11, 054105 (2017)
B. Preparation of the analogue fluids
The particulate analogue fluids used in this study were composed of polymethylmethacry-
late (PMMA) microbeads with a diameter of 10 lm, which have a similar dimension to human
erythrocytes. The microbeads were suspended at the same concentration as the RBCs (5% and
20% by weight) in Dx 40 and in the viscoelastic developed base fluid. The viscoelastic base
fluid was composed of 115 ppm (w/w) of xanthan gum (XG) diluted in Dx 40. More informa-
tion of the fluid preparation and composition can be found in the study by Calejo et al.41 A
summary of all the fluids used in this study can be found in Table I.
Deformability tests were also performed for 1% (w/w) of PMMA microbeads suspended in
the viscoelastic base fluid. For the deformability assessments, we have decided to use 1% of
microbeads and RBCs in order to obtain clear images of individual elements passing through
the hyperbolic contraction to reduce possible image analysis errors.
C. Fluid rheology
The rheological measurements in steady-state and small amplitude oscillatory shear
(SAOS) flows were carried out by means of a stress-controlled rheometer (Bohlin CVO,
Malvern, Worcestershire, UK) using a 55 mm diameter cone-plate geometry with a gap of
30 lm. According to Mezger,42 a gap size 5 times larger than the biggest particle size is recom-
mended, and in a more restrictive way, Schramm43 recommends a gap size 3 times larger than
the diameter of the particle. Hence, the gap (30 lm) used in this study was three times larger
than the dimensions of the particles (more details can be found in the supplementary material).
Steady shear flow curves were obtained in a range of shear rates of 1 _c=s1 10000. The
SAOS measurements were performed in a frequency range of 0.01 to 100 rad/s, after identifying
the range of amplitudes in which the elastic moduli (G0 and G00) behaved linearly. All the meas-
urements were carried out at 20 C, and at least three replicates in each measurement were
made in order to corroborate the reproducibility.
The large amplitude oscillatory shear (LAOS) measurements were performed with a plate-
plate geometry with a diameter of 50 mm and a gap of 1 mm, using a stress-controlled rheome-
ter from Anton Paar model Physica MCR-301. The measurements were achieved using a direct
strain oscillation module (DSO) by applying a strain amplitude sweep at an imposed frequency
of 0.1 and 1 rad/s. From the LAOS tests, we quantified the minimum and large strain elastic
moduli, G0M and G0L, respectively, and the minimum-rate and large-rate dynamic viscosities
g0M and g0L by following the framework described by Ewoldt et al.44 We also quantified the
TABLE I. Working fluids used in the experimental studies and the corresponding mass concentrations used (cell concentra-
tions by volume and PMMA microbeads by weight).
Suspended medium Dx 40
Blood fluids Dx 40 þ RBCs (5%)
Dx 40 þ Glucose-rich RBCs (5%)
Dx 40 þ RBCs (20%)
Dx 40 þ Glucose-rich RBCs (20%)
Two-phase fluids with a constant viscosity continuous phase Dx 40 þ PMMA (5%)
Dx 40 þ PMMA (20%)
Two-phase fluids with a variable viscosity continuous phase Dx 40 þ 115 ppm XG þ PMMA (5%)
Dx 40 þ 115 ppm XG þ PMMA (20%)
Deformability assessments Dx 40 þ RBCs (1%)
Dx 40 þ Glucose-rich RBCs (1%)
Dx 40 þ 115 ppm XG þ PMMA (1%)
054105-4 Pinho et al. Biomicrofluidics 11, 054105 (2017)
strain-stiffening (S) and shear-thickening (T) ratios measured using the same framework
(MITlaos version 2.2 beta for Matlab, freeware provided by [email protected]).44 These nonlin-
ear viscoelastic properties are well described in the literature by Ewoldt et al.,44,45 L€auger and
Stettin,46 and Sousa et al.47
D. Experimental setup and image analysis
1. Deformability analysis
Deformability tests were carried out for healthy and glucose-rich RBCs to observe and
quantify the influence of the high glucose concentration on the RBC deformability by using a
high-sensitivity microfluidic tool having a microchannel with a hyperbolic-shaped contraction.9
Similar measurements were also carried out for the analogue fluid, i.e., PMMA microbeads at
1% (w/w) suspended in the viscoelastic base fluid. These latter results were compared with data
obtained from the healthy and glucose-rich RBCs deformability tests.
To observe and analyse the behavior of the RBCs and the particulate analogue viscoelas-
tic fluid flowing through the PDMS microchannels with hyperbolic-shaped contractions, a
high-speed video microscopy system was used. This system consists of an inverted micro-
scope (IX71, Olympus, Japan) combined with a high-speed camera (Fastcam SA3, Photron,
USA). The PDMS microchannel was placed and fixed on the plate of the microscope, and the
flow rate of the working fluids was kept constant by means of a syringe pump (PHD Ultra,
Harvard Apparatus, USA) and a 1 ml syringe (Terumo, Japan). For the deformability experi-
ments, flow rates of 1 and 3 ll/min were used, and the images were captured at a rate of 3000
frames/s with a shutter speed ratio of 1/75 000 to minimize the dragging of the cells and
microbeads at high flow rates. All the flow experiments were performed at room temperature
(T¼ 20 6 2 C).
The recorded images were evaluated by means of the image analysis software ImageJ
(NIH)48 and by a similar procedure to that used by Pinho et al.8 and Rodrigues et al.19 Figure
1 shows the geometry of the microchannel and the regions where the deformation measure-
ments were taken for all the cells and microbeads. All regions were 200 lm long (see Fig. 1),
and the major (Major) and minor (Minor) dimensions of the cells or PMMA particles passing
through the regions were measured and their deformation index (DI) was calculated by using
the following equation:
DI ¼ Major Minorð Þ Major þ Minorð Þ : (1)
More detailed information about the cell deformability behavior when flowing through
a hyperbolic-shaped contraction microchannel can be found in the study by Yaginuma
et al.17
FIG. 1. Representation of the hyperbolic-shaped contraction followed by a sudden expansion used to assess the deformabil-
ity measurements with the selected Regions 1, 2, and 3.
054105-5 Pinho et al. Biomicrofluidics 11, 054105 (2017)
For these experiments, the same high-speed video microscopy system, described in the subsec-
tion above, was used. For this case, the flow rates of 1, 20, and 100ll/min were applied, and the
images were captured by the high speed camera at a frame rate of 2000 frames/s and a shutter speed
ratio of 1/10 000. All the flow experiments were performed at room temperature (T¼ 20 6 2 C).
The recorded image sequences were evaluated using ImageJ software using the function “Z proj-
ect”. The image analysis result is presented by a single image having a region brighter than the back-
ground, corresponding to the CFL region (Fig. 2). The intensity level distinction was used to identify
the CFL thickness since the position of the RBCs and of PMMA microbead core in the microchannel
is relatively clear. More details about this image analysis can be found in the study by Pinho et al.8
Note that in the CFL thickness analyses, we have only measured downstream of the con-
traction since for all the fluids studied in this work, we have observed that the CFL thickness
upstream is residual, as can be seen in Fig. 2.
E. Microchannels characterization
For the deformability analysis, a longer hyperbolic microchannel (MDI) having dimensions
of 396 15 15 lm, for the width (WuÞ, minimum contraction width ðWcÞ, and depth, respec-
tively, was used, cf. Table II. This high sensitivity microchannel is composed of a single hyper-
bolic contraction with a large Hencky strain of 3.3 followed by a sudden expansion. This large
Hencky strain not only generates a linear increase in the velocity but also allows an accurate
way to measure the deformation index (DI) of the cells and microbeads without tumbling or
rotational motions that are promoted by shear.
The CFL evaluation was performed downstream of contractions in different PDMS micro-
channels, designed to impose a large extensional flow deformation. In particular, one micro-
channel with a single hyperbolic-shaped contraction (M1) and a hyperbolic microchannel having
a sequence of 10 (M10) hyperbolic-shaped contractions, as shown in Table II, were used. The
evaluation is performed for the suspended healthy and glucose-rich RBCs and for the particu-
late viscoelastic fluids. The geometry of the hyperbolic contractions in the microchannels, M1
and M10, has the same dimensions with inlet and outlet widths of 398 lm (WuÞ, a length of
140 lm (L1Þ; and a minimum contraction width of 50 lm ðWcÞ, corresponding to a Hencky
strain of 2 (eH ¼ ln Wu=Wcð ÞÞ. The depth of all tested microchannels was 100 lm. Table II
summarizes the microchannels used and several important parameters.
A microchannel with an abrupt contraction followed by a triangular expansion was also used,
denoted as microchannel MT, to performed CFL measurements as it presents higher Hencky strain value
than the tested hyperbolic geometries, while maintaining the same aspect ratio (AR ¼ depth=Wu).
Most of the blood microchannel flow studies in the literature pertain to shear flow condi-
tions1–3,8,15 usually not taking into account the elastic nature of blood. However, in a more
FIG. 2. Example of the image analysis result and scheme of the CFL measurements performed downstream of the hyper-
bolic contraction, here for the working fluid containing PMMA microbeads.
054105-6 Pinho et al. Biomicrofluidics 11, 054105 (2017)
realistic approach to the microcirculation flow properties, the elastic behavior should also be
considered as it is significantly enhanced at small scale flows. Thus, we have decided to perform
flow experiments within hyperbolic-shaped contractions in which extensional elasticity properties
are significantly enhanced in particular for geometries with Hencky strains higher than 2. In addi-
tion, increasing the number of consecutive hyperbolic contractions and abrupt expansions enhan-
ces fluid stretching and provides an efficient way to investigate the elastic properties of fluids
with large extensional viscosities,36,49 as done in this work. Moreover, at these small dimensions,
the elasticity of the fluid can be assessed while limiting inertial effects. The microchannel MT
has an abrupt contraction followed by a smooth triangular expansion and also has a high Henchy
strain of 3 promoting the CFL formation downstream of the contraction, and as Rodrigues et al.9
conclude in their study, eH 3 has a strong impact on the CFL thickness.
The microchannel dimensions were also close to the dimensions of human microvessels
and in particular the strong extensional deformations imposed by the sequence of contractions,
and expansions may represent the effect of irregular networks that can be found in in vivo microcirculation. With these geometries and with the working flow rates, we were able to
achieve Reynolds numbers (Re ¼ qUWu=g) in the range of 0.01–3.5 and shear rates that can be
found at in vivo blood flow in microvessels with dimensions from 80 lm up to 500 lm.50
III. RESULTS AND DISCUSSION
1. Deformability analysis
The main purpose of this work is to develop analogue fluids that mimic in vitro suspen-
sions of pathological RBCs in terms of rheological and in vitro flow behavior. The pathological
TABLE II. Microchannel dimensions and experimental parameters: Hencky strain (eHÞ, aspect ratio ARð Þ; and Reynolds
number (Re).
M10 398 50 1378 2 0.25
MT 400 20 380 3 0.25
054105-7 Pinho et al. Biomicrofluidics 11, 054105 (2017)
RBCs were obtained by incubation in a rich-glucose medium, with 100 mM of glucose, for
twelve hours at 37 C. To better analyse and quantify the damage caused in the RBC mem-
branes by the high glucose concentration, deformability tests were performed with healthy and
glucose-rich RBCs suspended in Dx 40, using the microchannel MDI. The DIs obtained for the
pathological RBCs were compared with DIs of the healthy RBCs that were used as a control
sample. In Fig. 3, it is possible to observe that the incubated RBCs present a smaller DI than
the healthy RBCs for regions 1 and 2. For instance, in Region 1, the healthy RBCs present DIs of 0.36 and 0.54, for the flow rates of 1 and 3 ll/min, respectively, whereas the glucose-rich
RBCs present DIs of 0.27 and 0.41. In contrast, in Region 3, the pathological RBCs are more
deformed than the healthy cells. Taking into account these results, the same test was performed
with the PMMA particles, 1% by weight, suspended in the viscoelastic fluid (Dx 40þ 115 ppm),
in order to analyse also their ability to elongate when passing through the small contraction.
Overall, the PMMA microbeads always have the smallest DIs regardless of the region; how-
ever, some small elongations in Region 1 followed by a fast recovery to their original shapes in
the expansion region were observed. All deformability experiments were performed in the same
microchannel, MDI, at a room temperature of 20 6 2 C.
With this protocol, a similar in vitro condition to the stage of in vivo hyperglycaemia has
been developed, as the deformability results obtained for Region 1 clearly show that the incu-
bated RBCs have increased their rigidity. These results corroborate similar findings in the
literature.25,27–29
In contrast, the results obtained for Region 3 show that glucose-rich RBCs have a larger
DIs when compared with the healthy RBCs. This could be explained by the difficulty of the
glucose-rich RBCs in recovering their original shape due to the reduction in their membrane
viscoelastic properties.24,25 The PMMA microbeads also show some deformation; however, a
quick recover is observed in Regions 2 and 3. Hence, by performing the deformability tests, we
have demonstrated that the less deformable cells correspond to pathological in vitro RBCs. In
addition, the PMMA microbeads have demonstrated some ability to deform in Region 1, and
qualitatively, they were the best approximation to the behavior of rigid RBCs passing through
the hyperbolic contraction. In this way, PMMA microbeads can be a reasonable option as the
cellular component of the proposed analogue fluids. Note that the particles used in the literature
for similar studies34,38,39,41 are rigid or semi-rigid and larger; this is an important issue regard-
ing the development of two-phase analogue fluids because it is extremely difficult to develop
deformable micron-sized particles in a considerable quantity to make blood analogue fluids. By
using a flow focusing technique, Munoz-Sanchez et al.51 were able to produce a limited amount
FIG. 3. Deformation index of healthy and glucose-rich RBCs suspended in Dx 40 at 1% (/) and PMMA microbeads also
at 1% (w/w) suspended in the viscoelastic base fluid, flowing through the three selected regions of the hyperbolic micro-
channel MDI. The error bars represent the mean standard deviation at 95%, and *P < 0:05 was considered to be statistically
significant. All deformability tests were performed in microchannel MDI, at a room temperature of 20 6 2 C.
054105-8 Pinho et al. Biomicrofluidics 11, 054105 (2017)
of PDMS microbeads with diameters below 10 lm and elasticity close to healthy RBCs.
However, they still have to overcome several challenges such as the time of production, the
microbead final concentration, and the process to dry the microbeads.
Therefore, and in spite of the shortcomings in the behavior of the PMMA microbeads, as
in Region 3, we consider that they remain a good approximation to the representation of patho-
logical RBCs, as their values in Region 1 demonstrate that they also present some ability to
deform.
2. Aggregation assessment
It is known that diabetes and in particular the hyperglycaemia stages reduce the cell
deformability, promote the aggregation of the erythrocytes, and as a consequence, increase the
blood viscosity.28,52 In addition, in pathological microenvironments, the increment of the aggre-
gation level and blood viscosity promote the increase in the CFL thickness due to cell interac-
tions and the enhanced cell migration away from the microchannel wall.22,23,26,53 One of the
main objectives of this study is to investigate in detail the CFL development for all the pro-
posed analogue fluids, and bearing in mind that the incubation at a high glucose concentration
tends to increase the RBC aggregation levels, we have also decided to observe under the micro-
scope the level of aggregation for each sample.
The healthy RBCs suspended in plasma may promote the development of microstructures
called rouleaux with an organized structure similar to stacks of coins,4 although in pathological
environments three-dimensional aggregates are more likely to occur.4,22 To investigate the level
of aggregation, the same amount of healthy and glucose-rich cells and PMMA microbeads was
suspended in plasma and the formation of small clusters for all the tested samples were
observed. For the case of the healthy cells, the formation of simple rouleaux (one-dimensional)
was clear, as shown in Fig. 4(a), whereas for glucose-rich RBCs, the presence of complex
three-dimensional disorganized agglomerate structures was observed as shown in Fig. 4(b). For
PMMA microbeads suspended in plasma, the formation of organized structures was clearly
identified [Fig. 4(c)].
In in vitro blood flow experiments, Dx 40 is frequently used to prevent the sedimentation,
clotting, and jamming of cells and also works as a non-aggregating medium, as is possible to
observe in Fig. 4(d) where suspended healthy RBCs in Dx 40 show negligible rouleaux forma-
tion when compared with the healthy cells suspended in plasma [Fig. 4(a)]. However, when
FIG. 4. Microscopy images of human RBCs and microbeads suspended in plasma: (a) healthy RBCs, objective 20; (b)
glucose-rich RBCs, objective 20; and (c) PMMA microbeads, objective 10. Microscopy images of cells and microbead
suspensions in Dx 40: (d) healthy RBCs, objective 20; (e) glucose-rich RBCs, objective 10; and (f) PMMA microbeads,
objective 10.
054105-9 Pinho et al. Biomicrofluidics 11, 054105 (2017)
glucose-rich RBCs were suspended in Dx 40, this medium was less efficient as a non-
aggregating medium. Figure 4(e) shows the presence of rather large aggregated structures, con-
firming previous findings of the formation of larger aggregates when high concentrations of glu-
cose are involved.28,29 For the case of the PMMA microbeads, we have also observed the crea-
tion of organized structures within the Dx 40 medium. Overall, the performed aggregation
assessment shows that the glucose-rich RBCs present more aggregation then the healthy RBCs
suspended in the same medium and the PMMA microbeads present similar behavior in both
media.
1. Steady shear rheology of the suspended cells
Figure 5 compares the viscosity curves for the suspension medium Dx 40 and human
healthy RBCs suspended in Dx 40 with 5% and 20% of hematocrit. It is possible to observe
that Dx 40 has a constant viscosity of 5.2 mPa.s; however, as will be shown later in Sec.
III B 3, Dx 40 has some elasticity, so it behaves as a Boger-like fluid.54 By adding 5% and 20%
of RBCs, a clear increment of the viscosities of both fluids is obvious in comparison to the
pure Dx 40. At low shear rates, the curves for 5% and 20% hematocrit exhibit shear-thinning
up to shear rates of the order of few hundred reciprocal seconds, after which the suspension
tends to a constant viscosity (Newtonian behavior). This weak shear-thinning for the cells sus-
pended fluids is mainly associated with the cell deformability and less so to the weak formation
of rouleaux at low shear rates, as we have demonstrated in the aggregation studies. By increas-
ing the shear rate, the structures tend to break down and the deformable cells align, and as a
result, the viscosity decreases.
By increasing the RBC rigidity (glucose-rich RBCs), a slight increment in the viscosity val-
ues for the 5% of Hct with respect to the healthy RBCs can be noted, especially at low shear
rates [Fig. 6(a)]. However, taking into account the standard deviation of the data, this variation
is of no significance. The same effect is observed for 20% of Hct, as shown in Fig. 6(b).
2. Steady shear rheology of the developed analogue fluids
Figure 7 plots the shear rheology of fluids associated with the semi-rigid PMMA microbeads.
The addition of 5% of PMMA microbeads to the continuous phase increases the viscosity of Dx
40, but the viscosity of the suspension remains constant. However, adding 115 ppm of xanthan
gum to Dx 40 is sufficient to introduce shear-thinning (filled rhombus), and by further adding 5%
or 20% (w/w) of PMMA microbeads (10 lm of diameter), there is a viscosity increase, accompa-
nied by a small reduction of shear-thinning intensity as also shown in the data of Fig. 7, although
FIG. 5. Viscosity curves for the Dx 40 and healthy human RBCs at 5% and 20% of Hct suspended in Dx 40. All the meas-
urements were performed at 20 C, and the error bars represent the mean standard deviation.
054105-10 Pinho et al. Biomicrofluidics 11, 054105 (2017)
some shear-thinning is still visible. Note that both base fluids exhibit elasticity, with the Dx
40þ 115 ppm XG being more elastic as will be shown below (subsection III B 3).
Figure 8 compares the viscosity of the suspended glucose-rich RBCs with those of the pro-
posed particulate viscoelastic blood analogue. A reasonable agreement is observed, especially
for the 5% suspension, but small differences can still be found at low shear rates. By increasing
FIG. 7. Viscosity curves for the blood analogue fluids. All the measurements were performed at 20 C, and the error bars
represent the mean standard deviation.
FIG. 6. Viscosity curves for the human healthy and glucose-rich RBCs with: (a) 5% of Hct and (b) 20% of Hct suspended
in Dx 40. The error bars represent the mean standard deviation.
FIG. 8. Viscosity curves for both pathological cell suspensions and developed particulate blood analogue: (a) 5% of
glucose-rich RBCs and microbeads and (b) 20% of glucose-rich RBCs and microbeads. All the measurements were per-
formed at 20 C, and the error bars represent the mean standard deviation.
054105-11 Pinho et al. Biomicrofluidics 11, 054105 (2017)
the concentration of microbeads to 20%, the difference between both curves [Fig. 8(b)]
increases slightly possibly due to different abilities of the PMMA microbeads and glucose-rich
RBCs to deform under flow (glucose-rich RBCs deform better than PMMA microbeads as
shown in Fig. 3).
3. Oscillatory shear rheology
a. Small amplitude oscillatory shear. To perform a quantitative analysis of the relative magni-
tude of elastic and viscous behaviors of the developed analogues, measurements of the loss (G00) and storage moduli (G0) were carried out within the viscoelastic linear range, at an amplitude of
0.1 Pa. By analysing Fig. 9, it is possible to observe that the storage and loss moduli increase
with frequency, with the variation being stronger for the elastic component (G0). All fluids are
typically more viscous than elastic, with G00 being consistently higher than G0 up to a frequency
between 10 and 100 rad/s, in which there seems to be a crossover of G0 versus x for many of the
plots. We can also conclude that the viscous and elastic influences come mostly from the continu-
ous phase, Dx 40 and Dx 40þ 115 ppm of XG. As expected, the continuous phase of the devel-
oped analogue fluids is more elastic and more viscous than Dx 40, as seen in Fig. 9 (a).
The instrument low-torque limit line (Gmin ¼ FsTmin=c0)55 and the line of instrument inertia
(G0 ¼ ðIFs=FcÞx2Þ55 were added to Figs. 9 and 10 for a better understanding of the equipment
limitations and validity of the measurements. The line of instrument inertia is a slope 2 dashed
line, marking the onset of inertia effects, and the data for water plotted in Fig. 9 clearly show
that the variation of its G0 with x is solely due to inertia, as expected for a Newtonian fluid.
Clearly, the data for all working fluids (so, excluding water) all lie above the instrument inertia
limit even if at high frequencies the difference is small. With regard to the low torque limit
line (horizontal dotted-dashed line) only for the less elastic fluids, a limited number of data
FIG. 9. Storage and loss moduli of all the working fluids developed in this work at 20 C as a function of the angular fre-
quency: (a) continuous phase, (b) suspensions of healthy and glucose-rich cells, at 5% concentration in Dx 40, (c) at 20%
of concentration, and (d) developed analogue fluids at both microbead concentrations. The error bars represent the mean
standard deviation. The low-torque limit line represents the limit of accuracy of the rheometer Bohlin CVO.
054105-12 Pinho et al. Biomicrofluidics 11, 054105 (2017)
points lie below the limit (between 1 and 3 data points of 12 data points for each fluid), but
this takes place only at low frequencies (x< 0.1 rad/s). Therefore, with the exception of these
few points, the data sets in these figures are in a range of reliable measurements.
With regard to the suspensions of healthy and glucose-rich cells, at 5% concentration in
Dx 40 [Fig. 9(b)], there are no significant differences, whereas for the 20% suspensions, we
observe that G00 is higher for the glucose-rich RBCs than for the healthy cells [seen in Fig.
9(c)], as we have also observed in the steady shear data of Fig. 8(b). For the developed ana-
logues, a larger increase in G00 is also observed [Fig. 9(d)].
Hence, these results indicate that all the working fluids while being essentially viscous pre-
sent a non-negligible elastic contribution. Since a major objective is the development of analogue
fluids of pathologic cell suspensions, Fig. 10 composes the corresponding SAOS data comparing
the rheological behavior of the analogue fluids and of the glucose-rich RBCs. In Fig. 10, the lines
of slopes 1 and 2 were added for a better clarification of the increasing ratios of G0 and G00, and
the instrument low-torque limit line and the G0 of water have also been added.
We clearly observe in Fig. 10 that the analogue fluids are always more elastic and more
viscous than the suspensions of glucose-rich RBCs; however, taking into account the error bars,
the differences are not very significant. In addition, at least in the linear viscoelastic region, G0
and G00 of the analogue fluids represent very well the corresponding properties of the pathologi-
cal suspensions.
b. Large amplitude oscillatory shear. Large amplitude oscillatory shear (LAOS) tests were
performed varying the amplitude of the imposed shear stress at two angular frequencies of 0.1
and 1 rad/s. Three different maximum deformations were selected corresponding to three differ-
ent shear stresses of 0.02, 0.06, and 0.14 Pa, for the lower angular frequency (x¼ 0.1 rad/s),
and 0.16, 0.40 and 1 Pa for x¼ 1 rad/s. Fig. 11 plots the Lissajous-Bowditch curves normalized
with the first harmonic. The LAOS tests allow a complete characterization of the fluids, but
only the analogue fluids could be measured since ethical issues prevent the use of blood in the
rheometer equipped with this test.
At the lower angular frequency of 0.1 rad/s, all data plotted in Figs. 11(a) and 11(b) col-
lapse onto single curve behavior in the normalized Lissajous-Bowditch plots, indicating an
essentially viscous behavior, since the area enclosed by the curves in the shear stress versus
shear rate plot is very small (plot on right hand-side).
The behavior starts to change at the higher angular frequency of 1 rad/s, as is clear from
the inspection of Figs. 11(c) and 11(d). Here, while the behavior at large deformation exhibits
the same pattern seen at lower frequency, at the two lowest deformations, data are more dis-
perse and the enclosed area on the stress-rate of strain plot suggest some elasticity, albeit small.
From these results and by using the framework MTIlaos described by Ewoldt et al.,44 we
have also quantified some other quantities listed in Table III. The minimum strain elastic
FIG. 10. Storage (G0) and loss modulus (G00) for the suspension of glucose-rich RBCs in comparison to the correspondent
analogue fluid at (a) 5% and (b) 20%. The error bars represent the mean standard deviation. The low-torque and inertia
limit lines represent the limit of accuracy of the rheometer Bohlin CVO.
054105-13 Pinho et al. Biomicrofluidics 11, 054105 (2017)
FIG. 11. Normalized Lissajous-Bowditch plots of shear stress as a function of strain and shear rates at two angular frequen-
cies for the developed blood analogues: PMMA microbeads suspended in Dx 40þ 115 ppm XG at (a) 5% of PMMA,
x¼ 0.1 rad/s; (b) 20% of PMMA, x¼ 0.1 rad/s; (c) 5% of PMMA, x¼ 1 rad/s; and (d) 20% of PMMA, x¼ 1 rad/s.
054105-14 Pinho et al. Biomicrofluidics 11, 054105 (2017)
modulus, G0M, is represented by the tangent modulus at zero strain, and the maximum strain
elastic modulus, G0L, is represented by the secant modulus at the maximum strain. To represent
the viscous nature of the viscoelastic analogue fluids, the minimum-rate and large-rate dynamic
viscosities g0M and g0L were also quantified. The strain-stiffening and shear-thickening ratio, S and T, respectively, were also measured using the same framework that requires raw stress and
strain data.
These results, for all analogue fluids, indicate a negative shear-thickening ratio, T< 0, char-
acterizing a shear-thinning behavior during the oscillation cycle. Also, g0M > g0L corresponds
again to an intra-cycle shear-thinning behavior, corroborating the data obtained with the steady
shear tests. Regarding the elastic variables, it is possible to observe an intra-cycle strain-soften-
ing (S< 0) behavior since G0M>G0L even though both are very small, confirming the essen-
tially viscous nature of the analogue fluids and of their continuous phase. These findings are in
agreement with the intra-cycle behavior observed in Fig. 10, by the normalized Lissajous-
Bowditch plots of the stress versus strain and stress versus shear rate.
C. Cell-free layer assessment
In this study, we have performed flow visualizations and measurements of the CFL thick-
ness for the suspended healthy and glucose-rich RBCs flowing in the three different PDMS
microchannels (M1, M10, and MT) (cf. Table II). Note that, for all the fluids studied, we have
observed that the CFL thickness upstream of the contractions is residual (Fig. 12), so we have
only measured the CFL thickness downstream of the contractions. Figure 12 shows flows of
5% and 20% of glucose-rich RBCs suspended in Dx 40 and of PMMA microbeads suspended
in the viscoelastic fluid, upstream the sequence of contractions of the microchannel M10. In Fig.
12, it is possible to observe that the CFLs are residual or inexistent.
TABLE III. Minimum (G0M) and large (G0L) elastic moduli, minimum (g0M) and large–rate (g0L) dynamic viscosities, and
the strain-stiffening (S) and shear-thickening ratio (T) at two angular frequencies. Data obtained following the MITlaos pro-
cedure of Ewoldt et al.44 for the developed analogue fluids and PMMA microbeads at 5% and 20% concentrations sus-
pended in the viscoelastic base fluid are given.
x (rad/s) Blood analogue (%) Strain (%) G0M (Pa) G0L (Pa) g0M (Pa s) g0L (Pa s) S T
0.1 5 16080 1.27 106 8.77 106 1.21 102 1.09 102 0.44 0.10
20 4328 1.15 104 1.08 104 3.18 102 2.97 102 0.06 0.07
1 5 12490 1.62 104 1.10 104 9.72 103 8.97 103 0.47 0.08
20 8824 2.26 104 1.74 105 1.33 102 1.24 102 11.97 0.07
FIG. 12. Flow visualizations upstream the contraction sequence of the microchannel M10 for glucose-rich RBCs in Dx 40
and suspensions of PMMA microbeads in a viscoelastic medium, both at 5% and 20%. The objective is 20.
054105-15 Pinho et al. Biomicrofluidics 11, 054105 (2017)
In addition, comparisons with the CFL formed by the flow of the particulate blood ana-
logues were also performed, and the results are presented in Fig. 13. For all tested fluids, it has
been observed that downstream of the contractions of the microchannels, there is a high propen-
sity for CFL formation. The results show that the CFL thickness is influenced by the flow rate,
the number of sequential hyperbolic elements and hematocrit; increment with the hyperbolic
elements and flow rate, and decrease with the hematocrit [see Figs. 13(a) and 13(b)] for all the
fluids. These results corroborate qualitatively the blood flow studies performed by other
authors.8,9,12,22 The CFL thickness for the simple geometries M1 and MT and for 20% of
healthy RBCs (closer to in vivo microcirculation environments) is also in good agreement with
the in vivo results of Kim et al.6 and Yamaguchi et al.12
Since the analogue suspensions based on the Boger-like fluid continuous phase (Dx40þ PMMA
microbeads) showed CFL thicknesses close to zero, i.e., very far from the behavior of healthy or
pathological RBCs, and its rheological properties are also quite different from those suspension of
cells, we have decided not to present the data of this fluid. In summary, Dx 40þ PMMA microbeads
are a poor blood analogue, representing the suspensions of RBCs, healthy or pathological.
For the blood sample containing glucose-rich RBCs, the CFL thickness is always larger
than for the healthy human RBCs for all the tested microchannels. Hence, these latter results
reinforce the idea that the incubation applied to the healthy RBCs has indeed introduced
FIG. 13. CFL thickness for the three suspensions: healthy and glucose-rich RBCs and the corresponding proposed analogue
for 5% (a) and for 20% (b) of hematocrit. All the measurements were performed at 20 C, and the error bars represent the
mean standard deviation.
054105-16 Pinho et al. Biomicrofluidics 11, 054105 (2017)
changes to the structure and hemodynamic behavior of RBCs. One possible explanation for the
observed CFL increase of the samples containing glucose-rich RBCs might be due to their ten-
dency to form agglomerates, as already shown in Fig. 4. In the microcirculation, the aggrega-
tion of cells and their interactions are important factors that influence the hemodynamic and
hemorheological behavior of blood. For instance, several authors have demonstrated that an
increase in the RBCs aggregation, normally found at pathological conditions, promotes a signif-
icant increase in the CFL thickness,22,23,26 due to a more prominent migration of RBCs and
their aggregates, further increasing the cell aggregation level.6 This tendency to form larger
agglomerates increases also the sedimentation especially within the syringe reservoir and supply
tubes, whereas the sedimentation in the microchannels themselves is negligible considering the
flow transit time, except for the lowest flow rate. This is so because the syringe volume allows
for a long experiment in which the sedimentation occurs, while the fluid awaits its entrance
into the microchannel, especially for agglomerates, contributing to the decrease in the glucose-
rich RBCs concentration and consequently a more intense cell migration away from the wall,
resulting in a thicker CFL. The sedimentation time of the RBC agglomerates and the transit
times are estimated in the supplementary material.
During the flow experiments with the developed viscoelastic particulate analogues, a migra-
tion of the particles towards the centreline and the formation of a CFL thicker than for the
glucose-rich RBCs were observed. This increment of the CFL thickness could be due to the
shear-thinning behavior56–59 of the suspending fluid and sedimentation effect that is higher for
the PMMA microbeads, inducing a further increment in the CFL thickness. The sedimentation
of the PMMA particles along the inlet tubes and especially in the syringe is higher than that of
the glucose-rich RBCs since they present a higher sedimentation ratio, as shown in the supple-
mentary material [Figs. 16(c) and 16(d)]. Also, analysing their transit times inside of the micro-
channels, for all the flow rates, we can conclude that individual PMMA particles or RBCs do
not show significant sedimentation. Nevertheless, a significant reduction of the sedimentation
time is obtained when agglomerates are formed, which are lower than the transit time of the
flow within the microchannels for some low flow rates, as discussed in the supplementary mate-
rial. These effects are bound to have an influence on the CFL measurements by increasing their
thickness and consequently the difference between analogue fluids and glucose-rich RBCs.
The shear-thinning behavior, as observed by Li et al.,56 also promotes the motion of semi-
rigid microbeads away from the walls under high inertia or elasticity forces since the analogue
fluids were shown to be slightly more elastic and viscous than the suspension of glucose-rich
RBCs as seen in the rheological data, and this supports a higher migration of the PMMA
microbeads to the centreline than for the glucose-rich RBCs. This non-Newtonian effect com-
bined with the elastic response associated with the extensional flow of the hyperbolic contrac-
tions enhances the migration of particles towards the low shear rate region, i.e., to the centre
plane of the microchannel. This phenomenon is consistent with other experiments by Karimi
et al.,58 Calejo et al.,41 and Loon et al.60 who found that the migration of spherical and semi-
rigid particles depends strongly on the elasticity of the carrier fluid, and the migration towards
the centreline tends to occur when smooth shear-thinning base fluids are used,41,57–59 as the vis-
coelastic base fluid used in this work suspended the PMMA microbeads.
In addition, the numerical study by Li et al.56 has demonstrated that the elastic effects
drive the particles towards the channel centreline. They also showed that when the elastic and
inertial forces are equilibrated in the flow, the particle migration tends to an equilibrium posi-
tion in the microchannel. In this study, these elastic effects were stronger in the microchannel
M10 due to the sequence of contractions that will allow us to achieve a more evident elastic
response of the fluid. As we can see in Figs. 13(a) and 13(b), the microchannel M10 exhibited
larger differences in the CFL thickness between the glucose-rich RBC suspensions and ana-
logue fluids due to a stronger migration of the microbeads to the middle axis. It is also clear
that we have observed a migration of particles to the middle axis of the microchannel, and a
good agreement between the results was obtained in the microchannels M1 and MT.
For instance, by increasing the flow rate up to 100 ll/min and consequently to a Reynolds
value larger than 1, we have also obtained a higher shear rate and shear stress, leading to
054105-17 Pinho et al. Biomicrofluidics 11, 054105 (2017)
FIG. 14. Flow visualizations downstream of the sudden expansion of the microchannel M10 at flow rates of (a) 20 ll/min
and (b) 100 ll/min, for 5% of healthy and glucose-rich RBCs (Re 2.63) and PMMA microbeads suspended in the visco-
elastic medium (Re 2.35).
054105-18 Pinho et al. Biomicrofluidics 11, 054105 (2017)
stronger elastic effects, and at these conditions, we have indeed observed higher migration of
the RBCs and PMMA microbeads towards the region of lower shear rate, i.e., toward the cen-
treline of the microchannel, where the fluid does not feel the shear-tinning effect and the vis-
cous forces overcome the elastic forces, resulting in a ticker CFL for the three microchannels at
this flow rate. In Fig. 14, we can observe the evolution of the CFL thickness for the 5% of sus-
pensions in all continuous phases at flow rates of 20 ll/min [see Fig. 14(a)] and 100 ll/min [see
Fig. 14(b)].
Although there is a CFL difference between the analogue fluids and the suspensions of
glucose-rich RBCs, even larger differences were found between the analogue fluids and the
healthy RBC suspension. So, we believe that at the current stage, the proposed fluid is a reli-
able analogue fluid able to mimic the in vitro CFL phenomenon formed downstream of the con-
tractions by the glucose-rich RBCs. In particular, for the microchannel M1, similar results were
obtained between the analogue fluids and the suspensions of glucose-rich RBCs.
IV. CONCLUSIONS
The complexity and difficulties of working with blood in in vitro conditions, in particular
with pathological blood samples, call for the development of blood analogue fluids that are able
to mimic microcirculation phenomena, such as CFL, cell aggregation, and cell interactions in
order to improve our understanding of blood cell dynamics in the microcirculation for healthy
and pathological conditions. In this study, in vitro two-phase viscoelastic fluids were developed
which are able to mimic the rheological and flow behavior of pathological RBCs suspended in
dextran 40 (Dx 40) at room temperature. The analogue fluids were based on Dx 40 and relied
on suspended semi-rigid PMMA microbeads, exhibiting viscoelastic behavior due to the addi-
tion of a small quantity of xanthan gum.
The pathological RBCs were obtained by an incubation of healthy RBCs in high concentra-
tion of glucose, representing the pathological stage of hyperglycaemia in diabetic complications,
and an analysis of their deformation index (DI) was performed, showing that the glucose-rich
RBCs have a smaller DI than the healthy RBCs. Similar deformability tests were performed
with PMMA microbeads, showing that they present some ability to deform when flowing
through the hyperbolic contraction, so we used these semi-rigid microbeads to represent the cel-
lular component of the proposed analogue fluids.
Viscosity curves and small amplitude oscillatory shear tests were performed for the suspen-
sion of healthy and glucose-rich RBCs and also for the developed analogue fluids. A good
agreement between the viscosity curves and the linear viscoelastic moduli (G0 and G00) was
obtained for the suspension of pathological RBCs and the corresponding analogue fluids.
Overall, the rheological tests showed that the working fluids used here have a viscous compo-
nent that prevails over the elastic behavior, but nevertheless a non-negligible elastic character is
exhibited at 20 C. Large oscillatory shear tests were also performed to obtain a more extended
characterization of the analogue fluids, and the results corroborated the shear-thinning behavior.
Additionally, the ability of the developed particulate viscoelastic fluids to present a similar
flow behavior in comparison to the glucose-rich RBCs flowing within microfluidic devices was
analysed, in particular the cell-free layer (CFL) thickness downstream of a contraction. The
glucose-rich RBCs presented a thicker CFL than the suspensions of healthy RBCs. It is likely
that this CFL increment is in part associated with the presence of some agglomerates, sedimen-
tation effects, and changes in their shape and internal viscosity. Regarding the CFL formed by
the analogue fluid, suspensions of PMMA microbeads at concentrations of 5 and 20%, we have
observed that their CFL thickness resembles closely that of the fluid containing pathological
RBCs suspended in Dx 40.
In contrast to the majority of the existing blood analogues, this is the first analogue fluid
able to approximate not only the flow behavior of glucose-rich RBCs in microchannels but also
the rheological properties of the suspensions of RBCs. Therefore, it can be concluded from this
work that the developed analogue is a promising fluid to perform in vitro blood studies and
consequently to improve our understanding regarding the flow behavior of pathological blood
054105-19 Pinho et al. Biomicrofluidics 11, 054105 (2017)
cells in the microcirculation and for this particular case to mimic the pathological stage of
hyperglycemia in diabetes.
See supplementary material for detail information about the rheological measurements per-
formed with the stress-controlled rheometer Bohlin CVO with the gap size of 30 lm and about
the sedimentation effect observed during the flow visualizations.
ACKNOWLEDGMENTS
The authors acknowledge the financial support provided by Fundac~ao para a Ciencia e a
Tecnologia (FCT), COMPETE, and FEDER through the Ph.D. scholarship SFRH/BD/89077/2012,
Grant IF/00148/2013, and project POCI-01–0145-FEDER-016861 (with associated reference PTDC/
QEQ-FTT/4287/2014) funded by COMPETE2020—Programa Operacional Competitividade e
Internacionalizac~ao (POCI) with the financial support of FCT/MTES through national funds
(PIDDAC) and by the project Nos. PTDC/EQU-FTT/118716/2010, EXPL/EMS-SIS/2215/2013,
UID/EMS/00532/2013, and UID/EMS/04077/2013.
1R. Lima, T. Ishikawa, Y. Imai, M. Takeda, S. Wada, and T. Yamaguchi, J. Biomech. 41(10), 2188–2196 (2008). 2R. Lima, T. Ishikawa, Y. Imai, M. Takeda, S. Wada, and T. Yamaguchi, Ann. Biomed. Eng. 37(8), 1546–1559 (2009). 3R. Lima, T. Ishikawa, Y. Imai, and T. Yamaguchi, in In Single and Two-Phase Flows on Chemical and Biomedical Engineering, edited by R. Dias, A. A. Martins, R. Lima, and T. M. Mata (Bentham Science, 2012), pp. 513–547.
4G. McHedlishvili and N. Maeda, Jpn. J. Physiol. 51(1), 19–30 (2001). 5S. Kim, R. L. Kong, A. S. Popel, M. Intaglietta, and P. C. Johnson, Microcirculation 13(3), 199–207 (2006). 6S. Kim, P. K. Ong, O. Yalcin, M. Intaglietta, and P. C. Johnson, Biorheology 46(3), 181–189 (2009). 7M. Faivre, M. Abkarian, K. Bickraj, and H. A. Stone, Biorheology 43(2), 147–159 (2006). 8D. Pinho, T. Yaginuma, and R. Lima, BioChip J. 7(4), 367–374 (2013). 9R. O. Rodrigues, R. Lopes, D. Pinho, A. I. Pereira, V. Garcia, S. Gassmann, P. C. Sousa, and R. Lima, BioChip J. 10(1), 9–15 (2016).
10F. Robin and L. Torsten, The viscosity of the blood in narrow capillary tubes (American Physiological Society), Vol. 96, pp. 562–568.
11M. Nobuji, S. Yoji, T. Junya, and T. Norihiko, Erythrocyte flow and elasticity of microvessels evaluated by marginal cell-free layer and flow resistance (American Journal of Physiology – Heart and Circulatory Physiology, 1996), Vol. 271, pp. H2454–H2461.
12S. Yamaguchi, T. Yamakawa, and H. Niimi, Biorheology 29(2–3), 251–260 (1991). 13P. Vennemann, K. T. Kiger, R. Lindken, B. C. Groenendijk, S. Stekelenburg-de Vos, T. L. ten Hagen, N. T. Ursem, R. E.
Poelmann, J. Westerweel, and B. P. Hierck, J. Biomech. 39(7), 1191–1200 (2006). 14S. Cho, S. S. Ye, H. L. Leo, and S. Kim, in Visualization and Simulation of Complex Flows in Biomedical Engineering
(Springer, 2014), pp. 89–100. 15D. Pinho, R. O. Rodrigues, V. Faustino, T. Yaginuma, J. Exposto, and R. Lima, J. Biomech. 49(11), 2293–2298 (2016). 16T. Siddhartha, Y. V. B. V. Kumar, P. Amit, S. J. Suhas, and A. Amit, J. Micromech. Microeng. 25(8), 083001 (2015). 17T. Yaginuma, M. Oliveira, R. Lima, T. Ishikawa, and T. Yamaguchi, Biomicrofluidics 7(5), 054110 (2013). 18R. O. Rodrigues, M. Ba~nobre-Lopez, J. Gallo, P. B. Tavares, A. M. T. Silva, R. Lima, and H. T. Gomes, J. Nanoparticle
Res. 18(7), 194 (2016). 19R. O. Rodrigues, D. Pinho, V. Faustino, and R. Lima, Biomed. Microdevices 17(6), 108 (2015). 20D. A. Fedosov and G. Gompper, Soft Matter 10(17), 2961–2970 (2014). 21D. A. Fedosov, J. Fornleitner, and G. Gompper, Phys. Rev. Lett. 108(2), 028104 (2012). 22P. K. Ong, S. Jain, B. Namgung, Y. I. Woo, and S. Kim, Microcirculation 18(7), 541–551 (2011). 23J. Zhang, P. C. Johnson, and A. S. Popel, Microvasc. Res. 77(3), 265–272 (2009). 24N. Babu and M. Singh, Clinical Hemorheology Microcirculation 31, 273–280 (2004). 25J. Viskupicova, D. Blaskovic, S. Galiniak, M. Soszynski, G. Bartosz, L. Horakova, and I. Sadowska-Bartosz, Redox Biol.
5, 381–387 (2015). 26P. K. Ong and S. Kim, Microcirculation 20(5), 440–453 (2013). 27S. Sehyun, K. Yun-Hee, H. Jian-Xun, K. Yu-Kyung, S. Jang-Soo, and S. Megha, Clinical Hemorheology
Microcirculation 36, 253–261 (2007). 28S. Shin, Y.-H. Ku, J.-S. Suh, and M. Singh, Clin. Hemorheol. Microcirc. 38(3), 153–161 (2008). 29B. Riquelme, P. Foresto, M. D’Arrigo, J. Valverde, and R. Rasia, J. Biochem. Biophys. Methods 62(2), 131–141 (2005). 30V. Deplano, Y. Knapp, L. Bailly, and E. Bertrand, J. Biomech. 47(6), 1262–1269 (2014). 31A. D. Anastasiou, A. S. Spyrogianni, K. C. Koskinas, G. D. Giannoglou, and S. V. Paras, Med. Eng. Phys. 34(2),
211–218 (2012). 32F. J. H. Gijsen, F. N. van de Vosse, and J. D. Janssen, J. Biomech. 32(6), 601–608 (1999). 33J. D. Gray, I. Owen, and M. P. Escudier, Exp. Fluids 43(4), 535–546 (2007). 34T. T. Nguyen, Y. Biadillah, R. Mongrain, J. Brunette, J. C. Tardif, and O. F. Bertrand, J. Biomech. Eng. 126(4), 529–535
(2004). 35G. Vlastos, D. Lerche, and B. Koch, Biorheology 34, 19–36 (1997).
054105-20 Pinho et al. Biomicrofluidics 11, 054105 (2017)
36P. C. Sousa, F. T. Pinho, M. S. N. Oliveira, and M. A. Alves, Biomicrofluidics 5(1), 014108 (2011). 37L. Campo-Dea~no, R. P. A. Dullens, D. G. A. L. Aarts, F. T. Pinho, and M. S. N. Oliveira, Biomicrofluidics 7(3), 034102
(2013). 38O. Maruyama, T. Yamane, N. Tsunemoto, M. Nishida, T. Tsutsui, and T. Jikuya, Artif. Organs 23(3), 274–279 (1999). 39O. Maruyama, T. Yamane, M. Nishida, A. Aouidef, T. Tsutsui, T. Jikuya, and T. Masuzawa, ASAIO J. 48(4), 365–373
(2002). 40T. T. Nguyen, R. Mongrain, S. Prakash, and J. C. Tardif, paper presented at the Canadian Design Engineering Network
Conference, Montreal, QC, Canada (2004). 41J. Calejo, D. Pinho, F. Galindo-Rosales, R. Lima, and L. Campo-Dea~no, Micromachines 7(1), 4 (2016). 42T. G. Mezger, The Rheology Handbook: For Users of Rotational and Oscillatory Rheometers (Vincentz Network GmbH
& Co KG, 2006). 43G. Schramm, A Practical Approach to Rheology and Rheometry. (Gebrueder HAAKE GmbH, Karlsruhe, 2006). 44R. H. Ewoldt, P. Winter, and G. H. McKinley, MITlaos version 2.1 Beta for MATLAB (Cambridge, MA, 2007). 45R. H. Ewoldt, A. Hosoi, and G. H. McKinley, J. Rheol. 52(6), 1427–1458 (2008). 46J. L€auger and H. Stettin, Rheol. Acta 49(9), 909–930 (2010). 47P. Sousa, J. Carneiro, R. Vaz, A. Cerejo, F. T. Pinho, M. Alves, and M. Oliveira, Biorheology 50, 269–282 (2013). 48M. D. Abramoff, P. J. Magalh~aes, and S. J. Ram, Biophotonics Int. 11, 36–42 (2004). 49P. C. Sousa, F. T. Pinho, M. S. N. Oliveira, and M. A. Alves, J. Non-Newtonian Fluid Mech. 165(11-12), 652–671
(2010). 50L. Waite and J. M. Fine, Applied Biofluid Mechanics (McGraw-Hill Education, 2007). 51B. N. Munoz-Sanchez, S. F. Silva, D. Pinho, E. J. Vega, and R. Lima, Biomicrofluidics 10(1), 014122 (2016). 52K. Tsukada, E. Sekizuka, C. Oshio, and H. Minamitani, Microvasc. Res. 61(3), 231–239 (2001). 53B. Namgung, H. Sakai, and S. Kim, Clin. Hemorheol. Microcirc. 61(3), 445–457 (2015). 54L. Campo-Dea~no, F. J. Galindo-Rosales, F. T. Pinho, M. A. Alves, and M. S. N. Oliveira, J. Non-Newtonian Fluid Mech.
166(21–22), 1286–1296 (2011). 55R. H. Ewoldt, M. T. Johnston, and L. M. Caretta, in Complex Fluids in Biological Systems: Experiment, Theory, and
Computation, edited by S. E. Spagnolie (Springer New York, New York, NY, 2015), pp. 207–241. 56G. Li, G. H. McKinley, and A. M. Ardekani, J. Fluid Mech. 785, 486–505 (2015). 57R. P. Chhabra, Bubbes, Drops and Particles in Non-Newtonian Fluids, 2nd ed. (CRC Press, 2006). 58A. Karimi, S. Yazdi, and A. M. Ardekani, Biomicrofluidics 7(2), 021501 (2013). 59M. A. Tehrani, J. Rheol. 40(6), 1057–1077 (1996). 60S. V. Loon, J. Fransaer, C. Clasen, and J. Vermant, J. Rheol. 58(1), 237–254 (2014).
054105-21 Pinho et al. Biomicrofluidics 11, 054105 (2017)
http://dx.doi.org/10.1063/1.3567888
http://dx.doi.org/10.1063/1.4804649
http://dx.doi.org/10.1046/j.1525-1594.1999.06316.x
http://dx.doi.org/10.1097/00002480-200207000-00007
http://dx.doi.org/10.3390/mi7010004
http://dx.doi.org/10.1122/1.2970095
http://dx.doi.org/10.1007/s00397-010-0450-0
http://dx.doi.org/10.1016/j.jnnfm.2010.03.005
http://dx.doi.org/10.1063/1.4943007
http://dx.doi.org/10.1006/mvre.2001.2307
http://dx.doi.org/10.3233/CH-141909
http://dx.doi.org/10.1016/j.jnnfm.2011.08.006
http://dx.doi.org/10.1017/jfm.2015.619
http://dx.doi.org/10.1063/1.4799787
http://dx.doi.org/10.1122/1.550773
http://dx.doi.org/10.1122/1.4853455
s1
s2
s2A
s2B
s2C
t1
s2D
s2D1
d1
f1
s2D2
s2E
f2
s3
s3A
s3A1
t2
f3
s3A2
f4
s3B
s3B1
s3B2
f5
f7
f6
f8
s3B3
s3B3a
f9
s3B3b
f10
f11
s3C
t3
f12
f13
f14
s4
s5
c1
c2
c3
c4
c5
c6
c7
c8
c9
c10
c11
c12
c13
c14
c15
c16
c17
c18
c19
c20
c21
c22
c23
c24
c25
c26
c27
c28
c29
c30
c31
c32
c33
c34
c35
c36
c37
c38
c39
c40
c41
c42
c43
c44
c45
c46
c47
c48
c49
c50
c51
c52
c53
c54
c55
c56
c57
c58
c59
c60